Next Article in Journal
Nonlinear Modelling and Control of a Power Smoothing System for a Novel Wave Energy Converter Prototype
Next Article in Special Issue
Converting Offshore Oil and Gas Infrastructures into Renewable Energy Generation Plants: An Economic and Technical Analysis of the Decommissioning Delay in the Brazilian Case
Previous Article in Journal
Coordination of Automobile Supply Chain Considering Relative Endurance Level under the Dual-Credit Policy
Previous Article in Special Issue
Modeling and Simulation of Multipumping Photovoltaic Irrigation Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimal Multi-Objective Power Scheduling of a Residential Microgrid Considering Renewable Sources and Demand Response Technique

by
Mahmoud M. Gamil
1,2,*,
Soichirou Ueda
1,
Akito Nakadomari
1,
Keifa Vamba Konneh
1,
Tomonobu Senjyu
1,
Ashraf M. Hemeida
3 and
Mohammed Elsayed Lotfy
2
1
Department of Electrical and Electronics Engineering, Faculty of Engineering, University of The Ryukyus, 1 Senbaru, Nishihara-cho, Nakagami 903-0213, Okinawa, Japan
2
Department of Electrical Power and Machines, Zagazig University, Zagazig 44519, Egypt
3
Department of Electrical Engineering, Faculty of Energy Engineering, Aswan University, Aswan 81528, Egypt
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(21), 13709; https://doi.org/10.3390/su142113709
Submission received: 13 September 2022 / Revised: 7 October 2022 / Accepted: 19 October 2022 / Published: 22 October 2022
(This article belongs to the Special Issue Advanced Renewable Energy for Sustainability)

Abstract

:
Microgrid optimization is one of the most promising solutions to power system issues and new city electrification. This paper presents a strategy for optimal power scheduling of a residential microgrid depending on renewable generating sources and hydrogen power. Five scenarios of the microgrid are introduced to show the effect of using biomass energy and a seawater electrolyzer on microgrid cost and CO 2 emissions. Time of use demand response is applied to reshape the electric load demand and decrease the dependence on grid power. The obtained results from the multi-objective optimization verify that biomass has a significant role in minimizing the cost and CO 2 emissions; the cost is decreased by 37.9% when comparing scenarios with and without biomass. Besides, the FC integration with seawater electrolyzer and tanks reduces the microgrid emissions by around 40%.

1. Introduction

1.1. Greenhouse Gas Emissions

The growth of devastating greenhouse gas emissions acts as one of the main challenges to the human race [1]. Global carbon emissions in 2014 exceeded 1.6 times their levels in the 1990s [2]. GHGs aided in the significant rise in global temperature, posing a threat to human health and many economies [3]. Following the Kyoto Protocol, several countries have taken measures to cut GHG emissions, especially CO 2 emissions. GHGs half reduction by 2050 is a global target to face temperature increase [4]. In the power system sector, transferring to renewable sources is the best solution to achieve a significant reduction in GHG emissions [5].

1.2. Renewable Energy

Renewable power is generated from continuously rejuvenated energy flows such as wind, solar, geothermal heat, tidal, etc. [6]. Renewable energy sources are anticipated to provide 80% of global energy needs [7]. Globally, renewable energy capacity today stands at 2195 GW. The RES sector offered around 10.3 million jobs (directly and indirectly), with investments of over USD 280 billion [8]. A total of 145 countries have implemented programs to promote sustainable energy technologies during the last few decades. Biomass is a sustainable resource that may supplement other renewable energy sources. When combined with carbon capture, it creates no emissions [9]. Gasification is a reasonable method of generating electricity from biomass [10]. The growing popularity of renewable energy sources is helping to develop new microgrids.

1.3. Microgrid

Microgrids might be a viable solution to the current energy problems. A microgrid is made up of a group of dispersed electric sources and interconnected loads. Their operational modes are classified as isolated or grid-connected modes. Isolated microgrids are small power networks that produce and manage their own electricity needs. Grid connectivity improves system dependability and allows electricity trading with other grids [11,12]. Hydrogen technology is currently preferred in microgrids to store electric power. Electrolyzers normally produce hydrogen while power is available, while fuel cells generate energy from hydrogen during periods of low power [13]. The seawater electrolysis process produces chemical substances with hydrogen such as NaClO. It can be sold and manufactured to create revenue for the system. NaClO is used in drinking water disinfection, bleaching, and removing stains from clothing [14].

1.4. Motivations

The appropriate microgrid’s sizing has a major role in minimizing overall system costs, lowering CO 2 emissions, and promoting community development. Investment in the power system sector is an essential target to reduce the burden on the economy and encourage business in the energy sector. Microgrid productivity can be boosted through power trading and hydrogen technology. CO 2 emissions from power generation cannot be reduced without a reliance on renewable energy. Energy management technologies are strongly encouraged as a means of improving microgrid dependability and lowering microgrid costs [15].

1.5. Demand Response

Demand response programs support power system operators with cost-effective and energy-saving alternatives [16]. They represent customers’ adjustments to their energy consumption in response to changes in energy prices or incentive payments intended to limit power use at periods when the system’s stability is endangered [17]; they also help in reshaping load patterns. DR techniques are divided into price-based and incentive-based programs. In the price-based type, prices control electricity consumption by increasing peak-load hour tariffs while lowering them during off-peak hours. Consumers are rewarded for lowering their load during significant times via incentive-based demand response (DR) programs [18].

1.6. Related Literature

Many researchers have discussed the optimal design and operation of hybrid microgrids, considering renewable sources and demand response techniques. The appropriate power scheduling is greatly influenced by the desired sizing objectives. Cagnano et al. [19] outlined various control mechanisms required to achieve cost-effective, efficient, and secure operation of microgrids and reviewed the present primary design behaviors. A comprehensive literature assessment of existing microgrid-sizing methodologies was discussed in reference [20]. Cost-based and non-cost-based strategies are the two main sizing methodologies. For optimum energy cost and power supply probability, the Grasshopper Optimization Algorithm (GOA) was utilized for optimal microgrid sizing in [21]. In selecting the best size of microgrid, the net cost, renewable portion, energy cost, grid pricing rates, and greenhouse gas emissions were considered in [22]. Ref. [23] employed an evolutionary technique for optimum sizing of distributed energy sources to reduce the capital and yearly operating expenses. In ref. [24], HRES cost and different load values were considered for optimal microgrid sizing using the improved hybrid optimization genetic algorithm. Microgrid reliability was improved by connecting microgrids to the electric grid. Ref. [25] introduced a strategy for designing grid-connected microgrids that enhances their dependability while also serving the load at a low cost. For effective energy management of a grid-connected microgrid that depends on renewable energy, ref. [26] utilized the modified bat algorithm (MBA). MFABC+, MFABC, particle swarm algorithms, and HOMER software were utilized for optimal microgrid sizing in [27]. Ref. [28] used a multi-objective feasibility enhanced particle swarm optimization algorithm to reduce the microgrids’ operational costs and increase the renewable power use. Ref. [29] proposed optimal microgrid allocation based on renewable energy sources, demand response schemes, storage systems, and EV charging stations. Ref. [30] introduced optimal energy scheduling of grid-connected microgrids considering renewable sources uncertainty. Authors in [31] introduced two-step scheduling for microgrids using Monte Carlo and particle swarm optimization, considering sources’ uncertainty by grouping sources into basic load and frequency-modulated sources. Ref. [32] proposed an optimal energy management system for DC-microgrid composed of four nanogrids with a single energy storage system (ESS). Ref. [33] used genetic optimization to develop a novel multi-control scheme that maximized the power from renewable sources while minimizing total harmonic distortion. The results were compared with other optimization techniques. Power trading can enhance the system’s profit while lowering the total system cost via system sizing and optimization [34]. Ref. [35] demonstrated the impact of renewable sources on CO 2 level reduction. Biomass energy has proven itself as a suitable alternative energy source in electric power production. It is now considered a renewable energy source [36]. It has zero emissions when the CO 2 capture technique is employed [9]. Biomass gasification is a cheap solution for off-grid rural communities to produce electricity [10]. Ref. [37] researched biomass energy challenges as well as its large-scale employment. Ref. [38] studied the energy productivity of rice straw as a biomass fuel; it explored the burning of rice straw for energy and the issues associated with it. Ref. [39] discussed the latest microgrid control and power management studies. Ref. [40] presented different strategies for managing extra microgrid power with cost and emissions reduction as goals through selling surplus power to the grid, energy storage systems, and extra electricity conversion to hydrogen. The production of electricity from fuel cells via hydrogen production from seawater is regarded as the most abundant energy resource [41]. Hydrogen created from seawater in the morning can be used in fuel cells at night in microgrids that rely on PV [42]. Ref. [43] highlighted the need to reduce peak electricity demand.
Grid-purchased electricity and fuel costs can be reduced by employing DR schemes [44]. Ref. [45] studied the demand response objectives and classifications for facing electric power sector challenges. Demand response programs provide a cheap alternative to infrastructure upgrades in residential microgrids [46]. The restrictions and aims of demand response for diverse systems are described in [47]. Three types of demand response were compared in [48]; time-of-use demand response presents an effective and feasible technique to deal with peak load period problems. The barriers of participation in demand response were discussed in [49]. Ref. [50] used single- and multi-period load models to estimate emergency and time-of-use demand response programs based on load elasticity principles. The elasticity principal was used in ref 5 to address demand response [51]. With the help of thermostatically controlled and price-responsive loads, ref. [52] employed deep reinforcement learning algorithms to manage the energy of a grid-connected microgrid.

1.7. Paper Contribution

Regarding the previous literature, this paper’s contributions are outlined as follows:
  • Multi-objective optimal power scheduling of a residential microgrid considering revenues and productivity maximization of the microgrid using seawater electrolyzer and biomass generation.
  • The effects of load shifting techniques on reducing maximum demand and the grid’s power consumption, microgrid configuration, and emissions.
  • Introducing a comparison between different configurations for system design to demonstrate the feasibility and productivity of the used technologies.

1.8. Paper Construction

The remainder of the paper is organized as follows: Section 2 describes the system, and Section 3 clarifies demand response techniques. Objective functions are explained in Section 4, and constraints are discussed in Section 5. The optimization technique is discussed in Section 6. Simulation results and discussion are introduced in Section 7. Section 8 contains the conclusion and future work.

2. Microgrid Modeling

Optimal power scheduling is essential in the establishment process of microgrids to reduce the cost, emissions, loss of power supply, and the number of required system components. The system being studied is a residential microgrid in the northern part of Egypt. The proposed microgrid aims to meet all of its energy requirements with minimum cost and the least possible CO 2 emissions. Five case studies are discussed in this research to show the feasible system configuration. Figure 1 shows the available power generation units for all studied systems. The operation of each case study is demonstrated in Figure 2. Multi-objective genetic algorithm is utilized for the microgrid optimization with the help of MATLAB software.

2.1. Photovoltaic (PV) Modeling

PV transforms solar energy into electricity [53]. PV power is determined by solar radiation. Equation (1) describes the PV output power. This research uses Egyptian solar radiation data for the simulation; Egypt is considered one of the countries that has a wealth of solar power with a sunlight period of around 3500 to 4500 h/year [54].
P p v ( t ) = η p v × S p v × G ( t ) G S t c
where η p v , S p v , I d ( t ) , and I d , S t c are the PV’s efficiency, the rated capacity, the incident solar radiation, and solar radiation at the stc, respectively [55].

2.2. Fuel Cell (FC) Modeling

Fuel cell (FC) uses hydrogen to generate DC power. It normally consists of an electrolyte and two terminals (anode and cathode). The chemical processes inside FC are described in the equations below [56].
H 2 2 H + + 2 e
0.5 O 2 + 2 e O 2
2 H + + 1 2 O 2 2 H 2 O + h e a t
The capital cost and operating and maintenance costs of the fuel cell are described in the following equations:
C a p F C = α F C × S F C
O M F C = β F C × S F C × j = 1 N 1 + μ F C 1 + i r j
where C a p F C , α F C , S F C , O M F C , β F C , μ F C , i r , and N are the capital cost of FC, the investment cost of FC, FC capacity, operating and maintenance cost, the annual operating and maintenance cost, the escalation rate, interest rate, and project lifetime, respectively.

2.3. Sea Water Electrolyzer Modeling

Water covers approximately 75% of the earth’s surface. This water is mostly salty. Electrolysis of seawater can be utilized to create hydrogen as well as useful chemical substances [57]. The chemical reactions inside the electrolyzer are described by the following equations [58,59]:
N a C l + H 2 O N a C l O + H 2
2 N a + + 2 H 2 O + 2 e 2 N a O H + H 2
2 N a O H + C l 2 2 N a C l O + H 2
The overall equation during the electrolysis process is
N a C l + H 2 O N a C l O + H 2
The most appropriate method of hydrogen storage is pressurized gas storage. This requires a compressor and a storage tank [60].

2.4. Electric Utility

The utility grid serves as a backup when the generated power is not enough to supply the microgrid’s load. It also adds selling power availability to the grid in case of excess generation.
The exported and imported powers at a given time can be described using the equations below:
P g r p : P g r i d ( t ) > 0 i n c a s e o f i m p o r t i n g p o w e r
P g r s : P g r i d ( t ) < 0 i n c a s e o f e x p o r t i n g p o w e r
The net grid cost is
C g r i d = C p P g r p C S P g r s
where C g r i d , C p , P g r p , C S , and P g r s are the grid cost, unit power purchasing price, purchased power, unit power selling price, and the sold power, respectively [61,62].

2.5. Biomass Modeling

Biomass gasification is the process of converting solid biowaste into a combustible gas mixture. It can be used as a source of heat or as a fuel in internal combustion engines to produce mechanical or electric power. The calorific value of biomass and the amount of biomass determine its power [63]:
P b i o = T o t a l b i o m a s s a v a i l a b l e ( Ton / yr ) × 1000 × C V b m × η b m 365 × 860 × O p e r a t i n g h / day
The operating and maintaining cost of a biogas system is divided into two components (fixed and variable costs), which vary based on the anticipated power and the amount of fuel used:
O M b g , n p v = θ 1 b g × P b i o j = 1 N 1 + μ b g 1 + i r j + θ 2 b g × P W b g y r × j = 1 N 1 + μ b g 1 + i r j
Equation (16) shows the amount of biogas fuel cost:
F b g , n p v = θ 3 b g × B F r y r × j = 1 N 1 + μ b g 1 + i r j
Equations (17) and (18) explain the capital and salvage cost values of the biomass unit:
C b g = γ b g × P b i o
S V b g , n p v = λ b g × P b i o × 1 + δ 1 + i r N
where C V b m , η b m , θ 1 b g , P b i o , and i r are the biomass calorific valve, the overall conversion efficiency, the annual fixed operation and maintenance cost (USD/kW/year), the power produced by biogas generator, and the interest rate, respectively. μ b g , θ 2 b g , P W b g y r , and θ 3 b g are the escalation rate, the variable operation and maintenance cost (USD/kWh), the annual working power of biogas generator (kWh/year), and the biomass fuel cost (USD/ton), respectively. B F r y r , γ b g , λ b g , and δ are the annual required biomass fuel (ton/year), the initial cost of biogas system (USD/kW), the resale price of the system (USD/kW), and the inflation rate, respectively [64].

3. Demand Response

Demand response is an energy management approach that shifts energy usage from peak hours to other periods. This research studied time-of-use demand response as a time-based type.

3.1. Time-of-Use (ToU) Demand Response

In time-of-use demand response, peak load periods have higher prices, whereas off-peak hours have reduced prices according to predefined prices. In this work, price elasticity models are employed to structure the TOU demand response.

Elasticity Model

The change in load demand as a result of price swings is the price elasticity of electrical demand E l [65].
E l = ρ o d o × d ρ
where ρ , d o , ρ o , and d are the electricity price, the initial load demand, the nominal price, and the load demand, respectively.
The cross elasticity E l ( i , j ) illustrates how demand varies over time as a result of price changes at different time periods [66].
E l ( i , j ) = ρ o ( j ) d o ( i ) × d ( i ) ρ ( j )
The customer benefits are depicted as [67]
S = B ( d ( i ) ) d ( i ) × ρ ( i )
where B ( d ( i ) ) denotes the revenue earned by the usage of electrical energy as follows [68]:
B ( d ( i ) ) = B o ( i ) + ρ o ( i ) [ d ( i ) d o ( i ) ] { 1 + d ( i ) d o ( i ) 2 E l ( i ) × d o ( i ) }
Demand response advantages are boosted by setting S d ( i ) to zero. As a result, the following is the consumer usage:
d ( i ) = d o ( i ) { 1 + E l ( i ) × ρ ( i ) ρ o ( i ) ρ o ( i ) }
when the cross elasticity is taken into consideration, the load demand is expressed as follows:
d ( i ) = d o ( i ) + i = 1 , i j 24 E l ( i , j ) × d o ( i ) ρ o ( j ) × [ ρ ( j ) ρ o ( j ) ]
The final load demand that fulfills the maximum gains of the customer’s usage during a 24 h period is demonstrated in the following equation [68]:
d ( i ) = d o ( i ) { 1 + E l ( i ) × [ ρ ( i ) ρ o ( i ) ] ρ o ( i ) + i = 1 , i j 24 E l ( i , j ) × [ ρ ( j ) ρ o ( j ) ] ρ o ( j ) }
The self and cross elasticities are set on the basis of prices and demand to represent the flexibility to change the load patterns by shifting a portion of load from one period to another. The values of self and cross elasticities are mentioned in Table 1 [50].

4. Objective Function

This study introduces technical, economic, and environmental objectives for power scheduling of a residential microgrid, including PV, WG, and plug-in-electric vehicles. The first objective is the minimization of load-generation mismatch. The second objective is the minimization of total system cost. The third objective is CO 2 emissions minimization. The following equations show the proposed objective functions.
F 1 = m i n : L o P S = m i n : P l ( t ) P g n ( t )
F 2 = m i n : c o s t = m i n : C p v + C F C + C e l e c t r o l y z e r + C g r i d + C B i o C R e v e n u e s
F 3 = m i n : ( C O 2 ) e m i s s i o n s
where P g n , C p v , C F C , C e l e c t r o l y z e r , C g r i d , C B i o , and C R e v e n u e s are the total generated power, PV cost, fuel cell cost, electrolyzer cost, grid cost, total biomass cost, and system’s revenue, respectively.

5. Constraints

Power balance constraints:
P p v ( t ) + P F C ( t ) + P b i o + P g r i d _ b u y ( t ) P g r i d _ s e l l ( t ) P e l e c t r o l y z e r = P l ( t )
Limits constraints:
0 < P p v < P p v , m a x
0 < P F C < P F C , m a x
P g r s , m a x < P g r i d < P g r p , m a x
P b i o m i n < P b i o < P b i o m a x
H 2 t a n k m i n H 2 t a n k H 2 t a n k m a x

6. Multi-Objective Genetic Algorithm (MOGA)

The multi-objective genetic algorithm is a meta-heuristic mechanism motivated by the natural selection technique, which is a part of larger classes of evolutionary algorithms. Genetic algorithms are widely used and biologically inspired by developers to produce high-quality optimization and search prospects, such as mutation, crossover, and selection.
MOGA utilizes a weighted sum of various objective functions in the selection stage and merges them into a scalar fitness function. The design characteristics of the various objective functions weights are not specified and are randomly changed through each selection. Thus, the search orientation in this algorithm is not fixed.
At each generation over the process of MOGA, an empirical series of Pareto optimal solutions are stored and updated. Furthermore, a certain number of solutions are picked at random from the series. Those solutions are considered elite individuals. The elite mechanism has the benefit of preserving the diversity of each population [69].
The block diagram of the proposed MOGA algorithm is shown in Figure 3 and described below:
  • Stage 1 (Initialization): generate an initial population.
  • Stage 2 (Evaluation): calculate the values of the objective functions for the created population.
  • Stage 3 (Selection): use random weights to determine each population’s fitness value; then, pick a pair of strings from the existing population.
  • Stage 4 (Crossover and Mutation): a crossover strategy is implemented for each chosen pair to produce a new population via the crossover process; after that, the mutation process is carried out.
  • Stage 5 (Elitist): delete some strings of created strings haphazardly and substitute them with elite strings picked at random from temporary Pareto optimal solutions.
  • Stage 6 (Termination): if the stopping requirement is not satisfied, go to Stage 2.
  • Stage 7 (Optimal Solution): the MOGA suggests the preferable options.

7. Results and Discussion

Microgrid sizing and optimization are important issues for power system developers. The system configuration could differ depending on the available technology and the design objectives or targets. In this study, five case studies are introduced. In all scenarios, solar power has the first priority to produce the required electrical energy. The first scenario consists of PV, FC, seawater electrolyzer, tank, and public grid. The second scenario introduces biomass power to the first scenario but without the dependence on public grid. The third scenario has the availability of biomass and grid-connection with FC and seawater electrolyzer to guarantee full system reliability. Scenario 4 is like the system of the first scenario but without sea water electrolyzer. Scenario 5 is the simplest one, with PV and the public grid as the only power sources. The simulation parameters are shown in Appendix A. The load and generation of each power unit for all studied case studies are displayed in Figure 4, Figure 5, Figure 6, Figure 7 and Figure 8.
The simulation results for the five scenarios confirm the following features:
  • By taking scenario 5 as a reference case study because it is the simplest system configuration with the minimum number of generating units, the FC integration with sea water electrolyzer and tanks reduces the system emissions by around 40% and slightly increases the cost by USD 0.093 million.
  • If the microgrid that uses FC does not produce its own H 2 , its cost is increased due to the cost of purchasing H 2 . As seen from scenario 4’s results, its cost is greater than scenario 1’s cost by USD 0.246 million; the CO 2 emissions are also higher by 150.6 kg/day.
  • Relying on biomass has a great impact on cost and emissions. If biomass energy is used instead of depending on public grid power, the cost is decreased by 37.9%, as noticed by comparing scenarios 1 and 2. If both biomass power and electric utility are utilized in a hybrid microgrid as in scenario 3, the system has the lowest total cost of USD 1.186 million due to selling power back to the grid. Systems with biomass have zero emissions as they replace grid power use.
  • The grid power share is decreased by using biomass energy and FC; it reached zero by integrating biomass energy units. It decreased by 6% and 10% when comparing scenario 5 with scenarios 4 and 1.
  • Scenario 5 is the worst system configuration; it emits the highest emissions of 954.095 kg of CO 2 per day. Scenario 4 has the highest system cost and a great amount of CO 2 emissions, with roughly about 722.356 kg/day.
  • By comparing the systems that have storage tanks (1, 2, and 3), scenario 1 has the largest storage tank capacity as it has the largest fuel cell capacity of 184 kW with a power share percentage of more than 10%, as clarified in Figure 9. Scenario 2 has the largest EL capacity of 701 kW as it produces more chemical substances and hydrogen to increase revenue and decrease the system cost, as there is no revenue from the grid in this scenario.
Figure 9 demonstrates the contribution of the kWh energy production share of each generating unit.
  • In all the studied scenarios, the PV capacity ranges from 830 kW to 1000 kW. It provides more than 70% of the required power; the remaining percentage comes from the public grid, biomass, or stored energy in FC.
  • The incorporation of biomass power decreases the dependence of FC; it reduces the FC energy share by around 50% by comparing scenario 1 and scenario 2.
  • In scenario 3, the grid-connectivity is considered as a revenue tool to sell extra power back to the grid. It is regarded as a semi-grid-connected microgrid. Using seawater electrolyzers ensures the generation of the system’s required H 2 .
  • In scenario 3, the integration of biomass with EL and tank reduces the dependence on fuel cell; the capacity of the fuel cell is reduced to 30 kW.
  • Systems with seawater electrolyzers have the lowest CO 2 emissions. They reach 571.752 kg/day for systems without biomass energy.
  • The electrolyzing process is regarded not only as a means of producing hydrogen but also as a means of increasing system income. System productivity can be increased by selling extra power back to the public grid and selling NACLO and extra H 2 produced from the electrolyzing process.
  • Scenario 1 has a total system cost of USD 3.672 million with 571.752 kg of CO 2 emissions per day. By introducing biomass, both the emissions and the cost are enhanced. The cost is reduced by USD 1.394 million in scenario 2 and by USD 2.486 million in scenario 3 when compared with scenario 1.
Table 2 demonstrates the obtained results for all the studied scenario without DR programs.
  • All systems create revenues of more than USD 4 million. Scenarios 1 and 2 have the highest revenues of USD 4.47 million and USD 4.34 million, respectively.
  • The revenues from selling extra power back to the grid are decreased by introducing biomass to the system, in addition to the increase in selling chemical products and hydrogen produced from the seawater electrolyzer, as in scenario 3.
  • Without seawater electrolyzers, the revenues are only from selling power back to the public grid or nearby microgrids.
Table 3 shows the systems’ revenue for each scenario. Figure 10 displays the amount of revenue from selling power back to the grid and revenues from selling chemical products and hydrogen that are produced through the electrolysis process.
  • Demand response programs reshape the load patterns by shifting a portion of off-peak load, which is usually at night, to other periods. Biomass unit capacity is reduced by applying demand response as it is always used at night; it is reduced by 12.5% and 12.9% for scenarios 2 and 3.
  • By applying demand response schemes, the load curve is modified; the peak load is reduced by 10.88%, as seen in Figure 11.
  • The microgrids’ overall cost and CO 2 emissions are decreased with time-of-use demand response.
  • In most studied cases, the total system cost is reduced by TOU-DR with different values within USD 0.356 million, as in scenario 5, and USD 0.277 million, as in scenario 2.
  • The CO 2 emissions are reduced by 63.658, 20.763, and 71.819 kg/day for scenarios 1, 4, and 5, respectively. The maximum value of the grid’s purchased power is decreased by 13.71%, 8.6%, and 10.88% for scenarios 1, 4, and 5, respectively.
  • For scenarios without biomass, the FC capacity is decreased by applying DR schemes, while it is slightly increased for scenarios with biomass units to manage the decrease in biomass capacity.
Table 4 shows the scenarios’ configuration, cost, and emissions with their participation in ToU-DR schemes.

8. Conclusions

Microgrid size optimization plays a critical role in lowering total system costs by avoiding needless investment in unused generation. This study introduces a power scheduling methodology for a grid-connected microgrid considering PV, biomass power, fuel cell, and seawater electrolyzer. CO 2 emissions reduction, cost, and avoiding power outages are the main targets in the multi-objective scheduling process. The seawater electrolysis process is not only beneficial for producing hydrogen; it also serves as an income source for the system by selling the produced chemical compounds throughout the process. Time-of-use demand response is employed in this study to modify the load demand distribution and maximize the utilization of renewable energy sources. The research findings confirm that ToU-DR reduces the maximum load demand by 10.88%. They also confirm that CO 2 emissions can be reduced to zero by introducing biomass and also reduced by 40% by integrating FC and seawater electrolyzers. Biomass has the ability to decrease the microgrid cost by USD 1.39 million, if it replaces the grid power. Moreover, the hybrid microgrid of biomass, grid, PV, FC with seawater electrolyzer, and hydrogen tank is the most economical configuration. Furthermore, microgrid productivity is increased by selling both extra power and produced chemical products; it reaches over USD 4.1 million in most studied scenarios. Studying the integration of other renewable sources such as wind generators and integrating electric vehicles into the studied system through vehicles-to-grid schemes with a deep analysis of the sensitivity to parameter fluctuations is the proposed future work to increase microgrid productivity and reduce GHG emissions.

Author Contributions

Conceptualization, M.M.G. and K.V.K.; Data curation, M.M.G. and K.V.K.; Formal analysis, M.M.G. and T.S.; Investigation, M.M.G., A.N. and A.M.H.; Methodology, M.M.G., T.S. and A.M.H.; Project administration, M.M.G. and M.E.L.; Resources, M.M.G., K.V.K. and M.E.L.; Software, M.M.G., S.U. and M.E.L.; Supervision, T.S.; Validation, M.M.G. and A.M.H.; Visualization, M.M.G. and A.N.; Writing—original draft, M.M.G.; Writing—review & editing, M.M.G. and T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

GHGSGreenhouse gas emissions
ESSEnergy storage system
HRESHybrid renewable energy systems
MBAModified bat algorithm
GOAGrasshopper optimization
DRDemand response
PVPhotovoltaic
η p v The PV’s efficiency
S p v The PV’s rated capacity
G ( t ) The incident solar radiation
G S t c Solar radiation at the stc
C a p F C The capital cost of FC
α F C The investment cost of FC
S F C FC capacity
O M F C Operating and maintenance cost of FC
β F C The annual operating and maintenance cost
μ F C The escalation rate
i r Interest rate
NThe project lifetime
C g r i d The grid cost
C p Unit power purchasing price
P g r p Purchased power
C S Unit power selling price
P g r s The sold power
C V b m The biomass calorific valve
η b m The biogas overall conversion efficiency
θ 1 b g The annual fixed operation and maintenance cost of biogas generator (USD/kW/year)
P b i o The power produced by biogas generator
i r The interest rate
μ b g The escalation rate
θ 2 b g The variable operation and maintenance cost of biogas generator (USD/kWh)
P W b g y r The annual working power of biogas generator (kWh/year)
θ 3 b g The biomass fuel cost (USD/ton)
B F r y r The annual required biomass fuel (ton/year)
γ b g The initial cost of biogas system (USD/kW)
λ b g The resale price of the system (USD/kW)
δ The inflation rate
FCFuel cell
ToUTime of use
ElPrice elasticity of electrical demand
ρ The electricity price
d o The initial load demand
ρ o The nominal price
dThe load demand
E l ( i , j ) The cross elasticity
B ( d ( i ) ) The customer benefits
P g n The total generated power
C p v PV cost
C F C Fuel cell cost
C e l e c t r o l y z e r Electrolyzer cost
C g r i d Grid cost
C B i o Total biomass cost
C R e v e n u e s System’s revenue
LoPSLoss of power supply probability
P l ( t ) Load power
P p v ( t ) PV power
P F C ( t ) FC power
P b i o Biomass power
P g r i d _ b u y ( t ) Grid’s purchased power
P g r i d _ s e l l ( t ) Grid’s sold power
P e l e c t r o l y z e r Electrolyzer power
P g r i d ( t ) Grid power

Appendix A

Table A1. Simulation Parameters.
Table A1. Simulation Parameters.
Component and Economic Specification
Discount rate (r)5%
Escalation rate7%
PV Module
Investment cost1690USD/kW
Maintenance26USD/kW/yr
(PV) reduction factor84%
lifetime25years
Biomass generator
Capital cost4500USD/kW
Operating and Maintenance0.03USD/kWh
Feedstock cost0.02USD/kWh
Calorific value14.5MJ·kg 1
Electrical conversion efficiency ( η b m )0.3
Fuel cell
Capital cost2000USD/kW
Operating and Maintenance100USD/kW/yr
Replacement cost1500USD/kW
Efficiency0.5
H 2 to kW0.6kWhNm 3
Electrolyzer
Capital cost1500USD/kW
Operating and Maintenance15USD/kW/yr
Replacement cost1500USD/kW
Efficiency0.9
kW to H 2 0.09Nm 3 kWh
Final hydrogen pressure20MPa
Tank
Capital cost500USD/kg
Operating and Maintenance5USD/kg/yr

References

  1. Amagai, K.; Takarada, T.; Funatsu, M.; Nezu, K. Development of low-CO2-emission vehicles and utilization of local renewable energy for the vitalization of rural areas in Japan. IATSS Res. 2014, 37, 81–88. [Google Scholar] [CrossRef] [Green Version]
  2. Deng, X.; Lv, T. Power system planning with increasing variable renewable energy: A review of optimization models. J. Clean. Prod. 2020, 246, 118962. [Google Scholar] [CrossRef]
  3. Dong, Y.; Shimada, K. Evolution from the renewable portfolio standards to feed-in tariff for the deployment of renewable energy in Japan. Renew. Energy 2017, 107, 590–596. [Google Scholar] [CrossRef]
  4. Yao, S.; Zhang, S.; Zhang, X. Renewable energy, carbon emission and economic growth: A revised environmental Kuznets Curve perspective. J. Clean. Prod. 2019, 235, 1338–1352. [Google Scholar] [CrossRef]
  5. Esteban, M.; Zhang, Q.; Utama, A. Estimation of the energy storage requirement of a future 100% renewable energy system in Japan. Energy Policy 2012, 47, 22–31. [Google Scholar] [CrossRef]
  6. Twidell, J.; Weir, T. Renewable Energy Resources; Routledge: London, UK, 2015. [Google Scholar]
  7. Panwar, N.; Kaushik, S.; Kothari, S. Role of renewable energy sources in environmental protection: A review. Renew. Sustain. Energy Rev. 2011, 15, 1513–1524. [Google Scholar] [CrossRef]
  8. Ullah, S.; Haidar, A.M.; Hoole, P.; Zen, H.; Ahfock, T. The current state of Distributed Renewable Generation, challenges of interconnection and opportunities for energy conversion based DC microgrids. J. Clean. Prod. 2020, 273, 122777. [Google Scholar] [CrossRef]
  9. Lehtveer, M.; Fridahl, M. Managing variable renewables with biomass in the European electricity system: Emission targets and investment preferences. Energy 2020, 213, 118786. [Google Scholar] [CrossRef]
  10. Jimenez, O.; Curbelo, A.; Suarez, Y. Biomass based gasifier for providing electricity and thermal energy to off-grid locations in Cuba. Conceptual design. Energy Sustain. Dev. 2012, 16, 98–102. [Google Scholar] [CrossRef]
  11. Stadler, M.; Cardoso, G.; Mashayekh, S.; Forget, T.; DeForest, N.; Agarwal, A.; Schönbein, A. Value streams in microgrids: A literature review. Appl. Energy 2016, 162, 980–989. [Google Scholar] [CrossRef]
  12. Masrur, H.; Senjyu, T.; Islam, M.R.; Kouzani, A.Z.; Mahmud, M.P. Resilience-oriented dispatch of microgrids considering grid interruptions. IEEE Trans. Appl. Supercond. 2021, 31, 1–5. [Google Scholar] [CrossRef]
  13. Siwal, S.S.; Thakur, S.; Zhang, Q.; Thakur, V.K. Electrocatalysts for electrooxidation of direct alcohol fuel cell: Chemistry and applications. Mater. Today Chem. 2019, 14, 100182. [Google Scholar] [CrossRef]
  14. Asokan, K.; Subramanian, K. Design of a tank electrolyser for in-situ generation of NaClO. In Proceedings of the World Congress on Engineering and Computer Science, San Francisco, CA, USA, 20–22 October 2009; Volume 1, pp. 139–142. [Google Scholar]
  15. Gad, Y.; Diab, H.; Abdelsalam, M.; Galal, Y. Smart Energy Management System of Environmentally Friendly Microgrid Based on Grasshopper Optimization Technique. Energies 2020, 13, 5000. [Google Scholar] [CrossRef]
  16. Talaat, M.; Alsayyari, A.S.; Alblawi, A.; Hatata, A. Hybrid-cloud-based data processing for power system monitoring in smart grids. Sustain. Cities Soc. 2020, 55, 102049. [Google Scholar] [CrossRef]
  17. Balijepalli, V.M.; Pradhan, V.; Khaparde, S.A.; Shereef, R. Review of demand response under smart grid paradigm. In Proceedings of the ISGT2011-India, Kollam, India, 1–3 December 2011; pp. 236–243. [Google Scholar]
  18. Gao, J.; Ma, Z.; Guo, F. The influence of demand response on wind-integrated power system considering participation of the demand side. Energy 2019, 178, 723–738. [Google Scholar] [CrossRef]
  19. Cagnano, A.; De Tuglie, E.; Mancarella, P. Microgrids: Overview and guidelines for practical implementations and operation. Appl. Energy 2020, 258, 114039. [Google Scholar] [CrossRef]
  20. Alsaidan, I.; Alanazi, A.; Gao, W.; Wu, H.; Khodaei, A. State-of-the-art in microgrid-integrated distributed energy storage sizing. Energies 2017, 10, 1421. [Google Scholar] [CrossRef] [Green Version]
  21. Bukar, A.L.; Tan, C.W.; Lau, K.Y. Optimal sizing of an autonomous photovoltaic/wind/battery/diesel generator microgrid using grasshopper optimization algorithm. Sol. Energy 2019, 188, 685–696. [Google Scholar] [CrossRef]
  22. Nurunnabi, M.; Roy, N.K.; Pota, H.R. Optimal sizing of grid-tied hybrid renewable energy systems considering inverter to PV ratio—A case study. J. Renew. Sustain. Energy 2019, 11, 013505. [Google Scholar] [CrossRef]
  23. Logenthiran, T.; Srinivasan, D. Optimal selection and sizing of distributed energy resources for distributed power systems. J. Renew. Sustain. Energy 2012, 4, 053119. [Google Scholar] [CrossRef]
  24. Saiprasad, N.; Kalam, A.; Zayegh, A. Triple bottom line analysis and optimum sizing of renewable energy using improved hybrid optimization employing the genetic algorithm: A case study from India. Energies 2019, 12, 349. [Google Scholar] [CrossRef] [Green Version]
  25. Ecike, D. Using microgrids featuring PV panels and batteries connected to the grid to improve the reliability of a low-voltage feeder in Kinshasa. Energy Procedia 2019, 159, 117–122. [Google Scholar] [CrossRef]
  26. Luo, L.; Abdulkareem, S.S.; Rezvani, A.; Miveh, M.R.; Samad, S.; Aljojo, N.; Pazhoohesh, M. Optimal scheduling of a renewable based microgrid considering photovoltaic system and battery energy storage under uncertainty. J. Energy Storage 2020, 28, 101306. [Google Scholar] [CrossRef]
  27. Singh, S.; Slowik, A.; Kanwar, N.; Meena, N.K. Techno-Economic Feasibility Analysis of Grid-Connected Microgrid Design by Using a Modified Multi-Strategy Fusion Artificial Bee Colony Algorithm. Energies 2021, 14, 190. [Google Scholar] [CrossRef]
  28. Li, S.; Shi, L.; Yao, Z. multi-objective optimal scheduling of microgrid considering distributed generation uncertainty. In Proceedings of the 10th Renewable Power Generation Conference (RPG 2021), Online, 14–15 October 2021; pp. 35–43. [Google Scholar]
  29. Barik, A.K.; Das, D.C. Integrated resource planning in sustainable energy-based distributed microgrids. Sustain. Energy Technol. Assess. 2021, 48, 101622. [Google Scholar] [CrossRef]
  30. Chen, H.; Gao, L.; Zhang, Z. Multi-objective optimal scheduling of a microgrid with uncertainties of renewable power generation considering user satisfaction. Int. J. Electr. Power Energy Syst. 2021, 131, 107142. [Google Scholar] [CrossRef]
  31. Yang, M.; Cui, Y.; Wang, J. Multi-Objective optimal scheduling of island microgrids considering the uncertainty of renewable energy output. Int. J. Electr. Power Energy Syst. 2023, 144, 108619. [Google Scholar] [CrossRef]
  32. Zakir, M.; Arshad, A.; Sher, H.A.; Lehtonen, M. An Optimal Power Management System Based on Load Demand and Resources Availability for PV Fed DC-Microgrid with Power-Sharing among Multiple Nanogrids. In Proceedings of the 2021 IEEE PES Innovative Smart Grid Technologies Europe (ISGT Europe), Espoo, Finland, 18–21 October 2021; pp. 1–5. [Google Scholar]
  33. Sibtain, D.; Murtaza, A.F.; Ahmed, N.; Sher, H.A.; Gulzar, M.M. Multi control adaptive fractional order PID control approach for PV/wind connected grid system. Int. Trans. Electr. Energy Syst. 2021, 31, e12809. [Google Scholar] [CrossRef]
  34. Sediqi, M.M.; Howlader, H.O.R.; Ibrahimi, A.M.; Danish, M.S.S.; Sabory, N.R.; Senjyu, T. Development of renewable energy resources in Afghanistan for economically optimized cross-border electricity trading. Aims Energy 2017, 5, 691–717. [Google Scholar] [CrossRef]
  35. Chen, Y.; Wang, Z.; Zhong, Z. CO2 emissions, economic growth, renewable and non-renewable energy production and foreign trade in China. Renew. Energy 2019, 131, 208–216. [Google Scholar] [CrossRef]
  36. Uddin, M.; Taweekun, J.; Techato, K.; Rahman, M.; Mofijur, M.; Rasul, M. Sustainable biomass as an alternative energy source: Bangladesh perspective. Energy Procedia 2019, 160, 648–654. [Google Scholar] [CrossRef]
  37. Kyriakopoulos, G.L.; Arabatzis, G.; Chalikias, M. Renewables exploitation for energy production and biomass use for electricity generation. AIMS Energy 2016, 4, 762–803. [Google Scholar] [CrossRef]
  38. Liu, Z.; Xu, A.; Long, B. Energy from combustion of rice straw: Status and challenges to China. Energy Power Eng. 2011, 3, 325. [Google Scholar] [CrossRef] [Green Version]
  39. Elmouatamid, A.; Ouladsine, R.; Bakhouya, M.; El Kamoun, N.; Khaidar, M.; Zine-Dine, K. Review of Control and Energy Management Approaches in Micro-Grid Systems. Energies 2021, 14, 168. [Google Scholar] [CrossRef]
  40. Tabar, V.S.; Abbasi, V. Energy management in microgrid with considering high penetration of renewable resources and surplus power generation problem. Energy 2019, 189, 116264. [Google Scholar] [CrossRef]
  41. Fukuzumi, S.; Lee, Y.M.; Nam, W. Fuel production from seawater and fuel cells using seawater. ChemSusChem 2017, 10, 4264–4276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Srisiriwat, A.; Pirom, W. Feasibility study of seawater electrolysis for photovoltaic/fuel cell hybrid power system for the coastal areas in Thailand. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: London, UK, 2017; Volume 241, p. 012041. [Google Scholar]
  43. Morris, P.; Vine, D.; Buys, L. Residential consumer perspectives of effective peak electricity demand reduction interventions as an approach for low carbon communities. AIMS Energy 2016, 4, 536–556. [Google Scholar] [CrossRef] [Green Version]
  44. Harsh, P.; Das, D. Energy management in microgrid using incentive-based demand response and reconfigured network considering uncertainties in renewable energy sources. Sustain. Energy Technol. Assess. 2021, 46, 101225. [Google Scholar] [CrossRef]
  45. Jordehi, A.R. Optimisation of demand response in electric power systems, a review. Renew. Sustain. Energy Rev. 2019, 103, 308–319. [Google Scholar] [CrossRef]
  46. Davarzani, S.; Pisica, I.; Taylor, G.A.; Munisami, K.J. Residential Demand Response Strategies and Applications in Active Distribution Network Management. Renew. Sustain. Energy Rev. 2020, 138, 110567. [Google Scholar] [CrossRef]
  47. Patnam, B.S.K.; Pindoriya, N.M. Demand response in consumer-Centric electricity market: Mathematical models and optimization problems. Electr. Power Syst. Res. 2020, 193, 106923. [Google Scholar] [CrossRef]
  48. Yan, X.; Ozturk, Y.; Hu, Z.; Song, Y. A review on price-driven residential demand response. Renew. Sustain. Energy Rev. 2018, 96, 411–419. [Google Scholar] [CrossRef]
  49. Cardoso, C.A.; Torriti, J.; Lorincz, M. Making demand side response happen: A review of barriers in commercial and public organisations. Energy Res. Soc. Sci. 2020, 64, 101443. [Google Scholar] [CrossRef]
  50. Aalami, H.; Yousefi, G.; Moghadam, M.P. Demand response model considering EDRP and TOU programs. In Proceedings of the 2008 IEEE/PES Transmission and Distribution Conference and Exposition, Bogota, Colombia, 13–15 August 2008; pp. 1–6. [Google Scholar]
  51. Jalili, H.; Sheikh-El-Eslami, M.K.; Moghaddam, M.P.; Siano, P. Modeling of demand response programs based on market elasticity concept. J. Ambient Intell. Humaniz. Comput. 2019, 10, 2265–2276. [Google Scholar] [CrossRef]
  52. Nakabi, T.A.; Toivanen, P. Deep reinforcement learning for energy management in a microgrid with flexible demand. Sustain. Energy Grids Netw. 2021, 25, 100413. [Google Scholar] [CrossRef]
  53. El Mentaly, L.; Amghar, A.; Sahsah, H. The prediction of the maximum power of PV modules associated with a static converter under different environmental conditions. Mater. Today Proc. 2020, 24, 125–129. [Google Scholar] [CrossRef]
  54. International Renewable Energy Agency. Renewable Energy Outlook: Egypt; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates, 2018. [Google Scholar]
  55. Gamil, M.M.; Senjyu, T.; Takahashi, H.; Hemeida, A.M.; Krishna, N.; Lotfy, M.E. Optimal multi-objective sizing of a residential microgrid in Egypt with different ToU demand response percentages. Sustain. Cities Soc. 2021, 75, 103293. [Google Scholar] [CrossRef]
  56. Guaitolini, S.V.M.; Yahyaoui, I.; Fardin, J.F.; Encarnação, L.F.; Tadeo, F. A review of fuel cell and energy cogeneration technologies. In Proceedings of the 2018 9th International Renewable Energy Congress (IREC), Hammamet, Tunisia, 20–22 March 2018; pp. 1–6. [Google Scholar]
  57. Abdel-Aal, H.; Zohdy, K.; Kareem, M.A. Hydrogen production using sea water electrolysis. Open Fuel Cells J. 2010, 3, 1–7. [Google Scholar] [CrossRef]
  58. Donado, K.; Navarro, L.; Quintero M, C.G.; Pardo, M. HYRES: A multi-objective optimization tool for proper configuration of renewable hybrid energy systems. Energies 2019, 13, 26. [Google Scholar] [CrossRef] [Green Version]
  59. Casson, L.W.; Bess, J.W. On-Site Sodium Hypochlorite Generation; Water Environment Federation: Alexandria, VA, USA, 2006; pp. 6335–6352. [Google Scholar]
  60. Amos, W.A. Costs of Storing and Transporting Hydrogen; Technical Report; National Renewable Energy Lab. (NREL): Golden, CO, USA, 1999. [Google Scholar]
  61. Li, Z.; Xu, Y. Optimal coordinated energy dispatch of a multi-energy microgrid in grid-connected and islanded modes. Appl. Energy 2018, 210, 974–986. [Google Scholar] [CrossRef]
  62. Hossain, M.A.; Pota, H.R.; Squartini, S.; Abdou, A.F. Modified PSO algorithm for real-time energy management in grid-connected microgrids. Renew. Energy 2019, 136, 746–757. [Google Scholar] [CrossRef]
  63. Singh, S.; Singh, M.; Kaushik, S.C. Feasibility study of an islanded microgrid in rural area consisting of PV, wind, biomass and battery energy storage system. Energy Convers. Manag. 2016, 128, 178–190. [Google Scholar] [CrossRef]
  64. Heydari, A.; Askarzadeh, A. Optimization of a biomass-based photovoltaic power plant for an off-grid application subject to loss of power supply probability concept. Appl. Energy 2016, 165, 601–611. [Google Scholar] [CrossRef]
  65. Kirschen, D.; Strbac, G. Fundamentals of Power System Economics; John Wiley & Sons Ltd.: Chichester, UK, 2004. [Google Scholar]
  66. Kiptoo, M.K.; Adewuyi, O.B.; Lotfy, M.E.; Senjyu, T.; Mandal, P.; Abdel-Akher, M. Multi-objective optimal capacity planning for 100% renewable energy-based microgrid incorporating cost of demand-side flexibility management. Appl. Sci. 2019, 9, 3855. [Google Scholar] [CrossRef] [Green Version]
  67. Aalami, H.; Moghaddam, M.P.; Yousefi, G. Demand response modeling considering interruptible/curtailable loads and capacity market programs. Appl. Energy 2010, 87, 243–250. [Google Scholar] [CrossRef]
  68. Conteh, A.; Lotfy, M.E.; Adewuyi, O.B.; Mandal, P.; Takahashi, H.; Senjyu, T. Demand Response Economic Assessment with the Integration of Renewable Energy for Developing Electricity Markets. Sustainability 2020, 12, 2653. [Google Scholar] [CrossRef]
  69. Murata, T.; Ishibuchi, H. MOGA: Multi-objective genetic algorithms. In Proceedings of the IEEE International Conference on Evolutionary Computation, Perth, Australia, 29 November–1 December 1995; IEEE: Piscataway, NJ, USA, 1995; pp. 289–294. [Google Scholar]
Figure 1. The schematic diagram of all the studied microgrid configurations.
Figure 1. The schematic diagram of all the studied microgrid configurations.
Sustainability 14 13709 g001
Figure 2. The operational flowchart of all the studied cases.
Figure 2. The operational flowchart of all the studied cases.
Sustainability 14 13709 g002
Figure 3. MOGA operation flow chart.
Figure 3. MOGA operation flow chart.
Sustainability 14 13709 g003
Figure 4. Load-generation mismatch of case 1.
Figure 4. Load-generation mismatch of case 1.
Sustainability 14 13709 g004
Figure 5. Load-generation mismatch of case 2.
Figure 5. Load-generation mismatch of case 2.
Sustainability 14 13709 g005
Figure 6. Load-generation mismatch of case 3.
Figure 6. Load-generation mismatch of case 3.
Sustainability 14 13709 g006
Figure 7. Load-generation mismatch of case 4.
Figure 7. Load-generation mismatch of case 4.
Sustainability 14 13709 g007
Figure 8. Load-generation mismatch of case 5.
Figure 8. Load-generation mismatch of case 5.
Sustainability 14 13709 g008
Figure 9. kWh energy share.
Figure 9. kWh energy share.
Sustainability 14 13709 g009
Figure 10. Systems’ revenues.
Figure 10. Systems’ revenues.
Sustainability 14 13709 g010
Figure 11. Load demand with and without demand response.
Figure 11. Load demand with and without demand response.
Sustainability 14 13709 g011
Table 1. Self and cross elasticity values.
Table 1. Self and cross elasticity values.
PeakOff-PeakLow
Peak−0.10.0160.012
Off-Peak0.008−0.10.01
Low0.0060.008−0.1
Table 2. Results for systems without DR.
Table 2. Results for systems without DR.
Case 1Case 2Case 3Case 4Case 5
Cost (million USD)3.6722.2781.1863.9183.579
CO 2 emissions (kg/day)571.75200722.356954.095
PV capacity (kW)986.553830.866901.435995.8501000
FC capacity (kW)1846630970
EL capacity (kW)29470133600
Tank capacity (kg)147.5221000
Maximum grid power (kW)229.81000273.310321.810
Biomass capacity (kW)0288.972306.86100
Table 3. Revenues.
Table 3. Revenues.
Case 1Case 2Case 3Case 4Case 5
Grid revenue2.10401.4694.2324.253
Electrolyzing process revenue2.3664.1822.87500
Table 4. Results of systems with DR.
Table 4. Results of systems with DR.
Case 1Case 2Case3Case4Case5
Cost (million USD)3.3312.0011.3883.5623.247
CO 2 emissions (kg/day)508.09400701.593882.276
PV capacity (kW)980.965932.368874.959963.199993.292
FC capacity (kW)1776839740
EL capacity (kW)28578316800
Tank capacity (kg)147238300
Maximum grid power (kW)198.29700249.797286.797
Biomass capacity (kW)0252.934267.41300
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gamil, M.M.; Ueda, S.; Nakadomari, A.; Konneh, K.V.; Senjyu, T.; Hemeida, A.M.; Lotfy, M.E. Optimal Multi-Objective Power Scheduling of a Residential Microgrid Considering Renewable Sources and Demand Response Technique. Sustainability 2022, 14, 13709. https://doi.org/10.3390/su142113709

AMA Style

Gamil MM, Ueda S, Nakadomari A, Konneh KV, Senjyu T, Hemeida AM, Lotfy ME. Optimal Multi-Objective Power Scheduling of a Residential Microgrid Considering Renewable Sources and Demand Response Technique. Sustainability. 2022; 14(21):13709. https://doi.org/10.3390/su142113709

Chicago/Turabian Style

Gamil, Mahmoud M., Soichirou Ueda, Akito Nakadomari, Keifa Vamba Konneh, Tomonobu Senjyu, Ashraf M. Hemeida, and Mohammed Elsayed Lotfy. 2022. "Optimal Multi-Objective Power Scheduling of a Residential Microgrid Considering Renewable Sources and Demand Response Technique" Sustainability 14, no. 21: 13709. https://doi.org/10.3390/su142113709

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop